THE VOLATILE and intravenous anesthetics modulate activity of a variety of voltage-gated and ligand-gated ion channels. 1However, the mechanism(s) by which one or more of these effects produces general anesthesia and the specific site(s) of action which produce this state are presently unknown. There is growing recognition of the need for integrative approaches in order to relate anesthetic effects on ion channels with systems level behavior. 2–4However, the difficulty of determining how modulation of ion channel activity can produce general anesthesia is compounded by the relative paucity of information as to what constitutes general anesthesia at the systems level. Here, neural models are used to link anesthetic modulation of ion channel activity with overall systems level behavior using computer simulation and the analysis of the mathematical structure of some of these models. 5,6This complements recent work focusing primarily on anesthetic modulation of network interactions in more abstract 7and more complex 8,9models.

It has been hypothesized that the in vitro  concentration-effect curve of the receptor responsible for general anesthesia should be similar to that seen in vivo  at the systems level, where attention is generally paid to the slope and midpoint of such curves. 10An important feature of the in vivo  concentration-effect curve that must be accounted for is its threshold-like behavior as anesthetic concentration is increased. Since anesthetic concentrations that lead to immobility lie well below the EC50of the relatively shallow concentration-effect curves for volatile anesthetic suppression of voltage-gated ion channel activity, the above hypothesis has been used to implicate the ligand-gated ion channels as the site of action for the volatile anesthetics. 10 

The ligand-gated ion channel thought to be most involved in the mechanism of action of the volatile anesthetics is the GABAAreceptor which, with greater certainty, is the site of action for the intravenous anesthetics thiopental, propofol, etomidate, and the neurosteroids. 11Yet, volatile anesthetic action may not be completely understood in terms of anesthetic modulation of GABAA. Knockout mice lacking a component of the GABAAreceptor are readily anesthetized with volatile anesthetics, 12,13although this may simply reflect that the component of the GABAAreceptor necessary for volatile anesthetic action is still present in the knockout. Furthermore, the nonhalogenated anesthetics cyclopropane and butane have little effect on the GABAAreceptor 14which may also be true for the gaseous anesthetic xenon. 15,16However, if voltage-gated ion channels are relevant in general anesthesia, it is necessary to appreciate how relatively modest anesthetic effects at the ion channel are integrated to produce general anesthesia.

A model which links anesthetic action at the ion channel with systems level effects requires specific hypotheses as to the nature of general anesthesia at the systems level. Clinically, general anesthesia consists of at least loss of consciousness, amnesia, immobility, and blunting of autonomic reflexes. For inhalational anesthetics, anesthetic potency is generally quantified by using immobility as the endpoint. 17However, other attributes of general anesthesia appear at anesthetic concentrations that differ from those that lead to immobility. 18–22Moreover, immobility requires much greater anesthetic concentrations in preparations which provide anesthetic to only the brain and not the spinal cord. 23,24Thus, anesthetic effects at the systems level may be multimodal and associated with particular regions of the central nervous system.

Quantifying anesthetic effects at the systems level has been limited to analysis of the electroencephalograph and functional imaging of brain activity. Use of the electroencephalograph to characterize anesthetic depth has been complicated because electroencephalographic signals are not invariant with respect to anesthetic agent, 25and are not readily associated with specific endpoints such as immobility, 26though some progress has been made. 27Nonetheless, several features of the electroencephalograph are generally observed. For the intravenous anesthetics (barbiturates, propofol, and etomidate) a rather characteristic pattern is observed. At lower, generally subanesthetic, concentrations there is an increase in the fundamental frequency of the raw electroencephalographic signal, followed by slowing and increasing amplitude of the electroencephalograph with increasing anesthetic depth and, ultimately, burst suppression and isoelectric behavior are observed. Administration of benzodiazepines leads to loss of the α rhythm which is replaced by a higher frequency β rhythm in the human electroencephalograph, 28but does not eliminate the γ oscillations present in hippocampal slices. 8The volatile anesthetics are generally associated with increased slowing as anesthetic concentration is increased, followed by burst suppression and isoelectric behavior. 29Exceptions to this include seizure activity associated with enflurane, 30and lack of burst suppression or isoelectric electroencephalograph with the usual clinical concentrations of halothane. 31Increased variability is seen in the electroencephalograph during low dose ethanol consumption, and detailed analysis of the electroencephalograph under these conditions suggests a loss in the nonlinear structure of the electroencephalographic signal which permits an increased level of randomness. 32Functional imaging of human brain activity reveals that subanesthetic doses of isoflurane increase regional brain activation during specific tasks. 33During general anesthesia, the relative metabolism of the basal forebrain, thalamus, limbic system, cerebellum, occiput and midbrain reticular formation has been noted to decrease, 34,35and decreases in cerebral metabolic activity have been linked to measures derived from electroencephalographic activity. 36Regional differences in cerebral metabolic activity have been invoked to support the hypothesis 35that general anesthetics produce unconsciousness by modulation of the reticular nucleus of the thalamus. Thus, regional oscillatory activity as reflected by the electroencephalograph may be an important systems level correlate of anesthetic action.

Systems neurobiology has focused increasingly on the presence and role of correlated oscillations in the neocortex and other brain regions. Certain types of oscillations have been hypothesized to solve the “binding problem” of systems neurophysiology, the problem of linking the activation of multiple brain regions by related information. 37A neocortical layer of fast-spiking interneurons connected by inhibitory GABAAand electrical synapses may be responsible for generating γ frequency oscillations (40–80 Hz), 38–40which are hypothesized to solve the binding problem and appear to play an important role in attention. 41,42Other types of oscillations occur at lower frequencies, and have been associated with the level of consciousness. 43–46For example, the spindle oscillations (7 to 14 Hz) that originate in the thalamus and appear with the onset of sleep or with barbiturate anesthesia 43,47are probably generated in the reticular nucleus of the thalamus with modulatory input from thalamocortical relay cells. 48Thus, through modulation of the voltage- and ligand-gated ion channels, anesthetics could interfere with neural oscillations and the neurophysiologic processes regulated by these oscillations.

A number of relatively simple computational neural models are now examined to help conceptualize the integrated effects of general anesthetics on the behavior of individual neurons and larger networks.

Computational models of single neurons and networks of neurons were modified to permit modulation of the voltage-gated and ligand-gated ion channels by volatile and intravenous anesthetics. The effects of volatile anesthetics on neural activity were then evaluated as a function of anesthetic concentration and other model parameters. The impact on model behavior was assessed by computer simulation and, where possible, examination of the mathematical structure of the models using analytic techniques.

Models

Four computational models of individual neurons and networks of neurons were examined. These were the Morris–Lecar model of oscillation in the barnacle muscle fiber, 49the Pinsky–Rinzel simplification 50,51of the 19-compartment model of the CA-3 hippocampal neuron by Traub et al. , 52the Golomb–Rinzel network model of the reticular nucleus of the thalamus, 53and the Wang–Buzsáki model of hippocampal interneurons. 54These models were selected because of their relative simplicity and to illustrate specific aspects of anesthetic action. The Morris–Lecar and Pinsky–Rinzel models were used to examine the behavior of individual neurons as voltage-gated channel activity is modulated by anesthetics. The Golomb–Rinzel and Wang–Buzsáki models were used to examine anesthetic effects on simple oscillating networks.

General Principles of Computational Neural Models

Each of the biophysical models presented is built around the formalism developed by Hodgkin and Huxley. 55Membrane potential Vmis obtained from the differential equation for membrane potential:where C is the membrane capacitance, Iappis the externally applied current, and Iionis the sum of the currents resulting from the ion fluxes through the membrane ion channels. Essentially, this equation is a statement that the sum of the currents across the cell membrane must equal the externally applied current. Individual ionic currents whose sum comprises Iionare modeled as a potential across a conductance. Thus, the individual currents Iiare expressed as where giis the conductance of the channel and Viis the potential due to the ion conducted by the i-th channel (reversal potential). The conductance is a highly nonlinear function of time and membrane potential. A general expression for giis where ḡiis the maximal conductance of the i-th channel and is proportional to the channel density, piand qirepresent gates which must both be in a permissive state for current to flow, and liand miare the corresponding exponents which can be thought of as the number of gates in one ion channel. Both piand qirange between 0 and 1 to represent the fraction of gates in the permissive position. For both piand qithis fraction is typically governed by an equation of the form where α(Vm) is the voltage-dependent rate constant governing the rate that the gate transitions from the nonpermissive to the permissive state and β(Vm) is the voltage-dependent rate constant governing the rate that the gate transitions from the permissive to the nonpermissive state. Typically, α(Vm) and β(Vm) are highly nonlinear functions of membrane potential that must be empirically determined for each channel. For some systems calcium concentration can be a particularly important determinant of α and β. Temperature dependence of the model can be introduced through the system of equations defined by Eq. 4. This requires the use of a function φi(T), such that where φi(T) is often of the form where T is the new temperature, T0is the temperature for which the system was originally configured, and Q10-iis the coefficient describing the increase in rate of the i-th process for a 10-degree increase in temperature. Thus, the rates, but not the steady state solutions of the system of equations defined by Eqs. 5 and 6will be altered by temperature.

The above system of differential equations is a general description of a single compartment that can be considered the fundamental building block for constructing biophysical neural models. Often, a single compartment is sufficient to capture the behavior of interest as in the original Hodgkin–Huxley model or the Morris–Lecar model described below. To capture more elaborate neural behavior, particularly when investigating an anatomically complex neuron with a heterogeneous distribution of ion channels, multiple compartments can be electrically coupled together as in the Pinsky–Rinzel model, described below, and the more complex models which motivated it. Networks of single or multiple-compartment neurons can be generated by coupling them with one or more excitatory or inhibitory synapses as with the Golomb–Rinzel and Wang–Buzsáki models, below. Typically, synaptic currents are modeled beginning with equations like those of Eqs. 2–4, in which the conductance is governed by the presynaptic potential and the kinetics of the specific neurotransmitter. The contribution of each neuron in the network to the synaptic current(s) must be considered, and the connection geometries governing neural interactions can range from all-to-all, to random, to highly specific.

Specific Computational Models

Barnacle giant muscle fiber.

The Morris–Lecar model ( Appendix A) 49is useful for demonstrating the effects on individual neural activity of anesthetic modulation of voltage-gated ion channels. Although rather simple, this model displays many behavioral modes seen in more complex neural models and, therefore, is often used to demonstrate these different modes of behavior. 56Moreover, its relative simplicity permits more analytical approaches. It consists of a single compartment containing a noninactivating high voltage activated (HVA) Ca2+channel, a K+channel, and a leak current.

Hippocampal CA3 neuron.

The hippocampus plays an important role in memory and, because its anatomy is well suited for slice preparations, has been studied extensively in the laboratory. In addition, it has been the motivation for multiple computational modeling studies of increasing complexity, where this complexity extends to more realistic descriptions of ion channel activity, dendritic anatomy, distribution of ion channels over the dendrites, and network interactions. 57,58The Pinsky–Rinzel 50,51Appendix B) reduction of Traub et al.’  s 19-compartment hippocampal neuron model 52is used to illustrate the consequences of volatile anesthetic modulation of ion channel activity in hippocampal neurons. The two-compartment Pinsky–Rinzel model consists of single somatic and dendritic compartments that are electrically coupled. The somatic compartment contains a fast Na+channel, a fast K+channel, and a leak current. The dendritic compartment contains an HVA Ca2+channel, two Ca2+activated potassium channels (BK and SK), and a leak current.

Thalamic reticular nucleus.

Although the thalamus is primarily known for its multiple nuclei relaying afferent sensory input, the reticular nucleus of the thalamus is thought to be the source of the spindle oscillations (7–14 Hz) seen during drowsiness, at sleep onset, and barbiturate anesthesia, and is the source of the electroencephalographic α rhythm (8–12 Hz). 43,47,48The behavior of this nucleus is thought to be modulated by excitatory input from thalamocortical neurons which, themselves, receive several types of inhibitory input from the neurons of the reticular nucleus. Computational models of this system vary in complexity with respect to the description of the ion channels of the individual neurons as well as the nature and complexity of the interactions of the reticular nucleus with thalamocortical neurons. 53,59,60A simplified model of the reticular nucleus of the thalamus was developed by Golomb and Rinzel ( Appendix C) 53in which membrane potential is controlled by a low voltage activated (LVA) Ca2+channel and a leak current, and neural interactions are mediated by GABAAinhibitory interactions. Since modulation of GABAAreceptor activity is thought to be the basis for the action of a number of anesthetics, this model provides the opportunity to computationally examine the effects of anesthetics on a physiologically important neural system that may be related to consciousness. The model in  Appendix Chas been modified slightly from the original to permit partial connectivity between neurons in the network.

Fast-spiking interneuron network.

As delineated in the Introduction, interactions of the neocortex may be regulated by a network of fast-spiking interneurons. These fast-spiking interneurons may also contribute to the behavior of the hippocampus. 61,62Wang and Buzsáki (see  Appendix D) 54have constructed a model of this network in which neural interactions are modulated by GABAAinhibitory interactions. Because the voltage-gated ion channels that define the properties of the individual neurons in this network are different from those of the reticular nucleus of the thalamus, the network properties also differ. This difference arises primarily because the output of the thalamic neurons described above is a burst controlled by Ca2+entry as opposed to spikes controlled by the dynamics of fast-sodium and fast-potassium channels.

Anesthetic Modulation of Ion Channel Activity

The volatile anesthetics are known to modulate both voltage-gated 63and ligand-gated 64ion channels. Of the voltage-gated ion channels, the effect on the Ca2+channels is most pronounced, 63although other types of receptors may play a role. 65–69Inhibitory effects on both the LVA and HVA Ca2+channels decrease Ca2+influx under general anesthesia. 63Effects on the T- (LVA), L- (HVA), and N- (HVA) type Ca2+channels have been demonstrated. 63,70–72The decrease in Ca2+current as volatile anesthetic concentration is increased appears to be relatively voltage insensitive. 63Consequently, as a first approximation which neglects transient effects upon activation of the channel, 71it is reasonable to model volatile anesthetic modulation of Ca2+channels as a decrease in the maximal Ca2+conductance (ḡCain Eqs. A.1, B.2b, and C.1) which varies with volatile anesthetic concentration. The extent that a channel is inhibited as anesthetic concentration is increased is frequently modeled with the Hill equation where [A] is the anesthetic concentration, n is the Hill coefficient, and EC50is the anesthetic concentration at which channel activity is inhibited by 50%. This leads to the following modification of Eq. 3:where, effectively, the maximal conductance is now modulated as a function of anesthetic concentration. Volatile anesthetic effects on ion channels other than the Ca2+channels do occur. However, the effects on the fast-sodium and fast-potassium channels occur at much higher concentrations than for the Ca2+channels. 63,65,66Meaningful volatile anesthetic effects on the Ca2+-activated potassium channels may occur at anesthetic concentrations close to the EC50for the Ca2+channels. 73–76 

A variety of intravenous and volatile anesthetic effects are known for the ligand-gated ion channels. Most of these involve the GABAAinhibitory channel that controls neural interactions in the Golomb–Rinzel model 53of the reticular nucleus of the thalamus and the Wang–Buzsáki 54model of the fast-spiking interneuronal network. The GABAAinhibitory channel is the only ligand-gated ion channel in these models.

The intravenous anesthetics such as the barbiturates, 77,78etomidate, 79propofol, 78,80and neurosteroids 81are all known to increase the inhibitory chloride current by prolonging the open time of the GABAAchannel when it is stimulated with GABA. 80,82At higher concentrations of the intravenous anesthetics a decrease in the chloride current through the GABAAchannel is observed for at least several classes of the intravenous anesthetics which are known to prolong GABAAopening. 83,84At somewhat higher concentrations the intravenous anesthetics listed above can activate chloride currents in the absence of GABA. 80,81,85,86The benzodiazepines, though exerting their action at the GABAAreceptor, differ from the intravenous anesthetics in that they appear to exert their effect by increasing the rate at which the GABAAchannel opens, 87and are also distinguished by their inability to directly induce chloride currents through the GABAAreceptor. The effects of volatile anesthetics on the GABAAchannel share several similarities with the intravenous anesthetics. 64Their effect is characterized by prolongation of channel open time. However, at somewhat higher volatile anesthetic concentrations, the amplitude of the chloride current in the GABAAchannel is also decreased.

The above description of anesthetic effects on the GABAAchannel leads directly to a set of modifications to the descriptions of GABAAdynamics in the above computational models. An increase in the open time of the GABAAchannel, as seen with the intravenous and volatile anesthetics, can be modeled by decreasing the rate constant which governs the rate of channel closure in the equations describing GABAAchannel dynamics (βsynin Eqs. C.3 and D.4of the Appendices). This decrease in the rate constant as a function of anesthetic concentration has been measured experimentally for propofol and thiopental, 78and isoflurane, enflurane, and halothane. 64An increase in the rate of GABAAchannel opening, as seen with the benzodiazepines, can be modeled by increasing the parameter that governs the rate at which the channel opens (αsynin Eqs. C.3 and D.4of the Appendices). A decrease in the amplitude of the inhibitory chloride current generated when the GABAAis open can be modeled by decreasing the synaptic conductance (ḡsynin Eqs. C.1 and D.1of the Appendices). For some anesthetics, 64,78detailed quantitative descriptions of how these parameters vary with anesthetic concentration have been obtained and can be incorporated in the mathematical description of the model.

Computer Simulation

The models presented in the Appendices were implemented with Mathematica 4.02.2 88and were solved with the Runge–Kutta 89integration option. Additional simulations were also performed with the neural simulation language Genesis 902.0 that uses the exponential Euler integration scheme, where step size was progressively decreased and the stability of solutions with respect to chosen step size confirmed. All simulations were carried out until steady state behavior was achieved, and results from the terminal portion of the simulations are usually presented.

Analytical Determination of Underlying Model Structure

Analytic approaches become less practical with increasing model complexity. Consequently, these are considered only for the Morris–Lecar model. The example given here follows a very typical approach for problems of this type, 6and was previously exemplified for the Morris–Lecar model. 56Here, this approach is adjusted to examine the alteration in model behavior as anesthetic concentration is varied.

Although more complex questions can be addressed, the most fundamental issue is when and how qualitative behavior changes as one or more parameters are varied. In this case, the parameter is, effectively, anesthetic concentration through its modulation of ḡCa(Eq. A.1in the Appendices). The first step is to find the equilibrium point(s) of the system by setting the derivatives of Eqs. 1 and 4to zero and solving for the variables Vmand pi. For each of the equilibrium points, the system is linearized and the stability of the linearized system examined. This is typically done by computing the eigenvalues of the matrix which characterizes the linearized system at the given equilibrium point. If the eigenvalues of the matrix are given as λi, this corresponds to modes xi(t) of system behavior in the vicinity of the equilibrium point of the form

formula

which indicates that the solution will tend to leave the vicinity of the equilibrium point for Re(λi) > 0, will head toward the equilibrium point for Re(λi) < 0, and exhibit a transition (bifurcation) in behavior for Re(λi) = 0, where Re(·) denotes the real portion of the possibly complex eigenvalue. Thus, by examining λias parameters of the system are varied important information can be obtained about the ability of given parameters to impact on fundamental aspects of system behavior. A typical behavioral change indicated by Re(λi) passing through zero would be a switch from quiescence (Re(λi) < 0) to sustained oscillatory behavior (Re(λi) > 0).

All analytic calculations were carried out using both the symbolic and numeric capabilities of Mathematica. The equilibrium points were obtained numerically, these results were substituted into the corresponding Jacobian matrix, obtained symbolically from the linearized system, and the eigenvalues of this matrix were then computed.

Volatile anesthetic modulation of Ca2+-current in the barnacle muscle fiber

When parameterized as in  Appendix A, the Morris–Lecar model produces sustained oscillations for injection currents from 33 to 42 μA/cm2. In figure 1, for a fixed injection current of 35 μA/cm2, the maximal Ca2+conductance (ḡCain Eq. A.1) is decreased, as might occur in the presence of a volatile anesthetic, until oscillations cease. Initially, as ḡCais decreased, there is a graded decrease in the frequency of the oscillations (not shown). As illustrated in figure 1, for the given injection current, only a 10.8% decrease in the Ca2+conductance is sufficient to prevent sustained oscillations, and the termination of oscillatory behavior occurs abruptly as ḡCais decreased. The loci of injection current and ḡCawhere oscillations cease is shown in figure 2A. Note that when ḡCais decreased from its baseline value, larger injection currents are required for sustained oscillations. Conversely, larger injections currents will hyperpolarize the neuron unless ḡCais decreased. The Morris–Lecar model was developed for a physiologic preparation maintained at 22°C (room temperature). Volatile anesthetic sensitivity of HVA calcium channels at room temperature is known, and demonstrates a midpoint of approximately 0.85 mm and a Hill coefficient of 1.5 for halothane. 63Using these values to parameterize Eq. 8, the data of figure 2Acan be used to construct the loci of volatile anesthetic concentration and injection current, shown in figure 2B, for which oscillations cease. Simultaneously incorporating volatile anesthetic sensitivity of the potassium channels by assuming that halothane suppression of potassium channel activity has a midpoint of between 3.0 63and 6.4 mm 66and a Hill coefficient of 1.5 10introduces no additional behavior. For the points shown in figure 2, the termination of oscillatory behavior as ḡCais decreased corresponds analytically to a change in sign of the real component of one of the corresponding eigenvalues from positive to negative, demonstrating a basis for the abruptness of the transition.

Fig. 1. Behavior of the Morris–Lecar barnacle muscle fiber model, described in  Appendix A, as Ca2+conductance is varied with a fixed injection current. (A ) The behavior of the model for the parameters of  Appendix Ais shown at the top for an injection current of 35 μA/cm2. (B ) When the Ca2+conductance is decreased from its baseline value of 4.00 mS/cm2to 3.58 mS/cm2(10.5%), sustained oscillations persist. (C ) However, decreasing the Ca2+conductance to 3.57 mS/cm2(10.8% from baseline) prevents the emergence of sustained oscillations, demonstrating a threshold effect for behavior as Ca2+conductance is modulated.

Fig. 1. Behavior of the Morris–Lecar barnacle muscle fiber model, described in  Appendix A, as Ca2+conductance is varied with a fixed injection current. (A ) The behavior of the model for the parameters of  Appendix Ais shown at the top for an injection current of 35 μA/cm2. (B ) When the Ca2+conductance is decreased from its baseline value of 4.00 mS/cm2to 3.58 mS/cm2(10.5%), sustained oscillations persist. (C ) However, decreasing the Ca2+conductance to 3.57 mS/cm2(10.8% from baseline) prevents the emergence of sustained oscillations, demonstrating a threshold effect for behavior as Ca2+conductance is modulated.

Close modal

Fig. 2. Locus of injection current and Ca2+conductance (A ) or volatile anesthetic concentration (B ) where sustained oscillation is not possible for the Morris–Lecar barnacle muscle fiber model. As described in the text, the plot in (B ) was created from the data in (A ) assuming that halothane decreases Ca2+conductance according to the Hill equation where an halothane concentration of 0.85 mm deceases Ca2+conductance by 50% and with a Hill coefficient of 1.5.

Fig. 2. Locus of injection current and Ca2+conductance (A ) or volatile anesthetic concentration (B ) where sustained oscillation is not possible for the Morris–Lecar barnacle muscle fiber model. As described in the text, the plot in (B ) was created from the data in (A ) assuming that halothane decreases Ca2+conductance according to the Hill equation where an halothane concentration of 0.85 mm deceases Ca2+conductance by 50% and with a Hill coefficient of 1.5.

Close modal

Volatile Anesthetic Modulation of Hippocampal CA3 Neurons

The response of the Pinsky–Rinzel model of the CA3 hippocampal neuron (see  Appendix B) to steady current injection to the somal compartment is known to vary qualitatively with the magnitude of the injection current, 50as summarized in figure 3. Low levels of injection current lead to a bursting pattern (fig. 3A) mediated by calcium entry. Higher currents lead to complex patterns of spiking and bursting due to interactions between the somal and dendritic compartments (fig. 3B). Even higher currents lead to a somatic spiking pattern (fig. 3C), whose rate increases with increasing somal current injection (not shown) until the current is so great that the neuron remains in the depolarized state and cannot generate spike activity.

Fig. 3. Representative behavior of Pinsky–Rinzel model of CA3 hippocampal neurons, described in  Appendix B, during current injection to the somatic compartment. A bursting pattern is present in (A ) for IS= 0.75 μA/cm2. Increasing the injection current to IS= 1.75 μA/cm2leads to a complex interaction between the somal and dendritic compartments and a correspondingly complex pattern of bursting and spiking (B ). Higher injection currents lead to a spiking pattern generated by the somal compartment, shown in (C ) for IS= 2.5 μA/cm2. For each case, a segment near the end of the panel on the left is shown in greater detail on the right.

Fig. 3. Representative behavior of Pinsky–Rinzel model of CA3 hippocampal neurons, described in  Appendix B, during current injection to the somatic compartment. A bursting pattern is present in (A ) for IS= 0.75 μA/cm2. Increasing the injection current to IS= 1.75 μA/cm2leads to a complex interaction between the somal and dendritic compartments and a correspondingly complex pattern of bursting and spiking (B ). Higher injection currents lead to a spiking pattern generated by the somal compartment, shown in (C ) for IS= 2.5 μA/cm2. For each case, a segment near the end of the panel on the left is shown in greater detail on the right.

Close modal

Halothane modulation of maximal Ca2+conductance (ḡCain Eq. B.2b) was modeled using the same scheme and parameters as above. Modulation of the bursting pattern of figure 3Aby halothane is shown in figure 4for halothane concentrations of 0.10, 0.20, and 0.33 mm. Although a small increase in burst duration and frequency is seen in figure 4Afor concentrations of 0.10 mm, most of the increase occurs between 0.15 mm and 0.2 mm (fig. 4B). Between 0.2 mm and 0.33 mm, irregular behavior composed of shorter bursts and spikes eventually gives way to the regular spiking behavior shown in figure 4C. For somal injection currents that produce the irregular pattern of figure 3B, halothane modulation of Ca2+conductance produces a spiking pattern as halothane concentration is increased (not shown). Once the somal injection current is high enough to produce pure spiking behavior as in figure 3C, halothane modulation of Ca2+-conductance has no impact on the pattern and frequency of the spike train.

Fig. 4. Modulation of behavior of Pinsky–Rinzel model of CA3 hippocampal neurons by increasing concentrations of halothane during application of the relatively low somal injection current (IS= 0.75 μA/cm2) used to generate Fig. 3A. It is assumed that halothane modulates Ca2+conductance by decreasing the peak Ca2+conductance according to the Hill equation where a halothane concentration of 0.85 mm deceases Ca2+conductance by 50%, and with a Hill coefficient of 1.5. A modest increase in burst frequency and duration is produced when halothane concentration is 0.10 mm (A ). This becomes more pronounced for a halothane concentration of 0.20 mm (B ). Sustained spike activity appears with halothane concentrations of 0.33 mm (C ) and greater (not shown). For each case, a segment near the end of the panel on the left is shown in greater detail on the right.

Fig. 4. Modulation of behavior of Pinsky–Rinzel model of CA3 hippocampal neurons by increasing concentrations of halothane during application of the relatively low somal injection current (IS= 0.75 μA/cm2) used to generate Fig. 3A. It is assumed that halothane modulates Ca2+conductance by decreasing the peak Ca2+conductance according to the Hill equation where a halothane concentration of 0.85 mm deceases Ca2+conductance by 50%, and with a Hill coefficient of 1.5. A modest increase in burst frequency and duration is produced when halothane concentration is 0.10 mm (A ). This becomes more pronounced for a halothane concentration of 0.20 mm (B ). Sustained spike activity appears with halothane concentrations of 0.33 mm (C ) and greater (not shown). For each case, a segment near the end of the panel on the left is shown in greater detail on the right.

Close modal

The alterations in burst duration and frequency occur as a direct result of decreases in the Ca2+current. Although a decrease in Ca2+current requires a smaller dendritic K+current to repolarize the neuron, the burst duration still increases for a range of volatile anesthetic concentrations because decreases in Ca2+accumulation delay production of sufficient dendritic Ca2+-dependent K+currents (primarily IK-Cof Eq. B.2a) to terminate the burst. The decreased rate of intracellular Ca2+accumulation is due to both the decrease in the amplitude of the Ca2+current and, indirectly, to the corresponding decrease in the membrane potential of the dendritic compartment, which then prevents full activation of the gating parameter of the Ca2+channel (s in Eqs. B.2b, B.7 and B.8). The decrease in Ca2+concentration achieved during a burst also contributes to the observed increase in burst frequency. A burst cannot be initiated until the dendritic K+current (primarily IK-AHPof Eq. B.2a) is sufficiently small, which occurs when the gating variable q for IK-AHPdecreases below a specific threshold. When the Ca2+current is decreased by a volatile anesthetic, the decreased Ca2+concentration decreases the activation variable αq(Eq. B.14a) and, hence, the maximal value of q achieved during a burst. Since the decay of q is relatively invariant with respect to membrane potential and Ca2+concentration (Eq. B.14b), a smaller maximal value of q will require a shorter interval to reach the threshold where burst initiation becomes possible, leading to an increase in burst frequency. At higher volatile anesthetic concentrations, the Ca2+conductance is diminished to the point where Ca2+mediated voltage spikes do not occur in the dendritic compartment. Burst activity becomes impossible and is supplanted by rapid somatic spiking, effectively switching the behavioral mode of the neuron.

Incorporating volatile anesthetic effects on the sodium and potassium channels in the somal compartment adds little new behavior. The suppression of sodium channel activity by halothane has been reported to have a midpoint that ranges from 2.0 66to 2.6 63mm, and that for the potassium channel from 3.0 63to 6.4 66mm. Following earlier work, a Hill coefficient of 1.5 is assumed. 10When the more sensitive value of 2.0 mm is used for the sodium channel, the anesthetic concentration at which regular spiking behavior emerges will increase by about 0.01 mm, but only if the EC50of the potassium channel is close to 6.4 mm. Once regular spiking appears, it persists with a decreasing amplitude as anesthetic concentration is increased to at least several times the value of the midpoint of the sodium channel sensitivity.

Anesthetic Modulation of Inhibitory Interactions in the Reticular Nucleus of the Thalamus

The behavior of the Golomb–Rinzel model of the reticular nucleus of the thalamus as inhibitory interactions are modulated (fig. 5) for a network where, on average, each neuron interacts with 80% of the other network neurons. Here, the constant governing the rate at which the GABAAchannels close is varied. This might occur in the presence of intravenous and volatile anesthetics that modulate GABAAchannel behavior by prolonging the time that the GABAAchannel remains open. Smaller rate constants correspond to more prolonged channel opening. Measurements of the rate constant βsynyield a value of 0.067 · ms−1, which decreases to about 0.014 · ms−1in the presence of 200 μm thiopental and 10 μm propofol. 78The volatile anesthetics decrease βsynfrom about 2.5 (halothane) to 5 (isoflurane) times its baseline value. 64,91As βsynis decreased and GABAAchannels remain open longer, both the frequency of the overall neural activity and the coherence of this activity are affected (figs. 5, 6). The increased coherence as βsynis decreased is reflected in the increased amplitude and decreased frequency of the average membrane potential, which should be interpreted as a field potential analogous to the electroencephalograph. In contrast to overall network behavior, the gross behavior of individual neurons is minimally affected by changes in βsyn, as seen in figure 6.

Fig. 5. Behavior of network model of the reticular nucleus of the thalamus, described in  Appendix C, as βsynis decreased, corresponding to a prolongation in the time that GABAAchannels remain open. The membrane potential for a single neuron within the network is shown in the left hand-hand column. The average of the membrane potential from all 100 neurons in the simulation is shown in the middle column and is the analog of the electroencephalograph for the model. A raster plot for all 100 neurons is shown in the right-hand column, and was generated by indicating when the membrane potential of a given neuron exceeds -55 mV. Simulations are shown for βsynequal to 0.08 (A ), 0.04 (B ), 0.01 (C ), and 0.005 · ms−1(D ).

Fig. 5. Behavior of network model of the reticular nucleus of the thalamus, described in  Appendix C, as βsynis decreased, corresponding to a prolongation in the time that GABAAchannels remain open. The membrane potential for a single neuron within the network is shown in the left hand-hand column. The average of the membrane potential from all 100 neurons in the simulation is shown in the middle column and is the analog of the electroencephalograph for the model. A raster plot for all 100 neurons is shown in the right-hand column, and was generated by indicating when the membrane potential of a given neuron exceeds -55 mV. Simulations are shown for βsynequal to 0.08 (A ), 0.04 (B ), 0.01 (C ), and 0.005 · ms−1(D ).

Close modal

Fig. 6. Frequency of network activity for simulation of the reticular nucleus of the thalamus as βsyn.is increased, corresponding to an increase in the rate of GABAAchannel closure. The fundamental frequency of the average membrane potential from all neurons in the network (filled circles) is shown along with the mean (± SD) of the fundamental frequency of the membrane potential of the individual neurons (open circles). As GABAAchannels remain open longer (smaller βsyn), the decrease in the frequency of the average membrane potential arises primary from modulation of the collective behavior of the network since the frequency of individual neural behavior is relatively unchanged.

Fig. 6. Frequency of network activity for simulation of the reticular nucleus of the thalamus as βsyn.is increased, corresponding to an increase in the rate of GABAAchannel closure. The fundamental frequency of the average membrane potential from all neurons in the network (filled circles) is shown along with the mean (± SD) of the fundamental frequency of the membrane potential of the individual neurons (open circles). As GABAAchannels remain open longer (smaller βsyn), the decrease in the frequency of the average membrane potential arises primary from modulation of the collective behavior of the network since the frequency of individual neural behavior is relatively unchanged.

Close modal

Decreases in maximal Ca2+conductance (ḡCaof Eq. C.1) of about 20–40%, as might occur in the presence of the volatile 63or intravenous anesthetics, 92permits network synchrony to occur at somewhat larger values of βsyn(not shown). The network behavior that is most prominent with decreased Ca2+conductance is the increase in the value of βsynwhere the fundamental frequency of the average membrane potential approaches that of the individual neurons, signifying relatively complete network synchrony. This occurs for βsyn=.01 · ms−1when ḡCa= .4 mS/cm2, and βsyn= .02 · ms−1when ḡCa= .3 mS/cm2. The decreased Ca2+current permits a relatively greater impact of the inhibitory currents on membrane potential, effectively increasing the coupling between neurons, permitting synchronization to occur at larger values of βsyn. Unlike the Pinsky–Rinzel model of the CA3 hippocampal neuron presented above, the relatively simple calcium dynamics of the Golomb–Rinzel model of the reticular nucleus of the thalamus precludes refinement of the calcium burst by delayed calcium entry. Consequently, the shape of the calcium burst and the burst frequency remain unaffected with the above decrements in Ca2+conductance.

Both the volatile 64,91and intravenous 83,84anesthetics may decrease the amplitude of the inhibitory currents produced by GABAAchannels. As might be anticipated, 53the above results are affected minimally by decreases of 10–20% in the conductance governing GABAAsynaptic interactions (ḡsynin Eq. C.1). Decreases of this magnitude might be observed at relatively high concentrations of the volatile anesthetics. 64An increase in the rate constant governing GABAAopening (αsynin Eq. C.3), as is hypothesized to occur in the presences of benzodiazepines, 87has little effect on the behavior of individual neurons or the entire network. This occurs because the onset of inhibitory currents is already very rapid when compared to the relatively slow time course of the individual neurons. Consequently, increasing αsyndoes little to alter the time course or intensity of inhibitory interactions.

Anesthetic Modulation of Inhibitory Interactions in a Fast-Spiking Interneuronal Network

As emphasized by Wang and Buzsáki 54and illustrated in figure 7, the synchrony of the fast-spiking interneuronal network depends critically on the rate constant governing the closure of the GABAAsynapses. However, in marked contrast to the network model of the reticular nucleus of the thalamus, reduction in the value of βsynleads to decreased  synchrony of the network. This is reflected by an increase in frequency and decrease in amplitude of the membrane potential when it is averaged over all neurons in the network as a network analog of the electroencephalograph. Importantly, this occurs for relatively modest reductions in βsyn, especially when compared to the reductions in βsynthat were necessary to achieve network synchrony in the model of the reticular nucleus of the thalamus. The reduction in synchrony with reduction of βsynwas seen for a range of values of ḡsyn. In contrast to the properties of the Golomb–Rinzel model depicted in figure 6, the frequencies at which individual neurons in the network oscillate decrease somewhat with decreasing βsyn.

Fig. 7. Behavior of fast-spiking interneuron model, described in  Appendix D, as βsynis decreased, corresponding to a prolongation in time that GABAAchannels remain open. The membrane potential for an individual neuron is shown in the left hand-hand column. The average membrane potential from all 100 neurons in the simulation is shown in the right-hand column and is the model analog of the electroencephalograph. Simulations are shown for βsynequal to 0.08 (A ) and 0.06 (B ) ms−1. The injection current for each neuron was chosen randomly from a Gaussian distribution with a mean of 1.0 μA/cm2and SD 0.03 μA/cm2.

Fig. 7. Behavior of fast-spiking interneuron model, described in  Appendix D, as βsynis decreased, corresponding to a prolongation in time that GABAAchannels remain open. The membrane potential for an individual neuron is shown in the left hand-hand column. The average membrane potential from all 100 neurons in the simulation is shown in the right-hand column and is the model analog of the electroencephalograph. Simulations are shown for βsynequal to 0.08 (A ) and 0.06 (B ) ms−1. The injection current for each neuron was chosen randomly from a Gaussian distribution with a mean of 1.0 μA/cm2and SD 0.03 μA/cm2.

Close modal

The sensitivity of the fast-spiking interneuronal network to alterations in the rate at which the GABAAchannels open, as might occur in the presence of the benzodiazepines, is shown in figure 8. An increase in the rate at which GABAAchannels open can lead to greater degrees of synchronous behavior. Although the values of αsynused to generate figure 8are somewhat smaller than that of 12 · ms−1, used to generate the simulations of figure 7, they demonstrate a sensitivity of the fast-spiking interneuron model to changes in the rate of GABAAchannel opening that are not present in the more slowly oscillating model of the reticular nucleus of the thalamus.

Fig. 8. Behavior of fast-spiking Interneuron model as αsynis increased, corresponding to an increase in the rate at which GABAAchannels open. The membrane potential for an individual neuron is shown in the left-hand column. The average membrane potential from all 100 neurons in the simulation is shown in the right-hand column and is the analog of the electroencephalograph for the model. Simulations are shown for αsynequal to 1 · ms−1(A ) and 6 · ms−1(B ). Note how increasing αsynleads to more synchronous behavior. Further increases in αsyndo not lead to any fundamentally different behavior from that shown here.

Fig. 8. Behavior of fast-spiking Interneuron model as αsynis increased, corresponding to an increase in the rate at which GABAAchannels open. The membrane potential for an individual neuron is shown in the left-hand column. The average membrane potential from all 100 neurons in the simulation is shown in the right-hand column and is the analog of the electroencephalograph for the model. Simulations are shown for αsynequal to 1 · ms−1(A ) and 6 · ms−1(B ). Note how increasing αsynleads to more synchronous behavior. Further increases in αsyndo not lead to any fundamentally different behavior from that shown here.

Close modal

Incorporating volatile anesthetic effects on Na+and K+conductance as described for the model of the hippocampal neuron has minimal impact on the behavior of individual neurons within the network. However, the decreased spike amplitude as anesthetic concentration approaches the EC50of the Na+channel leads to decreased inhibitory interactions and loss of network synchrony for halothane concentrations of about 1.5 mm.

To help conceptualize the integrated response to anesthetic modulation of voltage-gated and ligand-gated ion channels and the contribution of this integrated response to general anesthesia, four computational neural models were modified to incorporate known features of anesthetic action on ion channel behavior. Models were chosen for their parsimony and their ability to illustrate integrative aspects of anesthetic action in systems with features that may be relevant to general anesthesia.

Models and Their Limitations

Barnacle Giant Muscle Fiber.

The Morris–Lecar model of the barnacle giant muscle fiber, 49although of minimal clinical relevance, is an important model of calcium oscillations whose simplicity permits some degree of mathematical analysis which complements observations made from the simulations. The major finding with this model is that relatively modest decrements in maximal Ca2+conductance, as might occur in the presence of the volatile anesthetics, can terminate the oscillations. Importantly, the Ca2+conductance and, therefore, the anesthetic concentration at which neural quiescence emerges is a function of the level of stimulation provided by the amplitude of the injection current, consistent with the observation that general anesthesia is a function of the level of stimulation. An examination of the underlying mathematical structure of the model reveals not just a change in the nature of the oscillation, but the emergence of a perfectly stable system that is no longer capable of oscillation.

Although it is known that volatile anesthetics can suppress Ca2+current, obtaining model behavior as a function of the concentration of a particular volatile anesthetic, as was done in figure 2for the Morris–Lecar model, is still somewhat imprecise. First, the kinetics of the volatile anesthetic's interaction with the Ca2+channel are not well-established, nor is the effect of the volatile anesthetic on the kinetics of the Ca2+current. Second, the full complement of necessary experimental data are unavailable. Minimally, this would include measurements of anesthetic modulation of ion channel activity for the same type of preparation as the model, obtained at the same temperature as the experimental data on which the model is based, and with sufficient measurements to describe channel behavior over the full range of anesthetic concentrations. The Morris–Lecar model was developed for a preparation of the barnacle giant muscle fiber at room temperature (22°C). Data obtained at room temperature from a different type of cell demonstrates values of 0.85 mm for the EC50, and 1.5 for the Hill coefficient (n in Eq. 7), 63which were used along with Eq. 7to estimate volatile anesthetic effects for other anesthetic concentrations. Even though there is some imprecision in relating volatile anesthetic concentration to decreases in calcium current, this does not alter the fundamental observations of threshold-like behavior occurring for decreases in maximal Ca2+conductance well below 50%, and that the threshold varies with the level of stimulation current.

Hippocampal CA3 Neuron.

The Pinsky–Rinzel 50model of the CA3 hippocampal neuron was examined with the same computational approach used to examine the Morris–Lecar model of the barnacle giant muscle fiber. To determine the impact of volatile anesthetics on model behavior, the model incorporated the same quantitative relationship between anesthetic concentration and maximal Ca2+conductance (ḡCain Eqs. B.2b and B.8) used above in the Morris–Lecar model. Volatile anesthetic modulation of Ca2+current leads to an increase in burst frequency, a lengthening of burst duration and, finally, a switch from bursting to spiking behavior as anesthetic concentration is increased (fig. 4). As detailed in the Results, all of these effects occur in the model as a direct result of decreases in the Ca2+current and the subsequent decrease in intracellular calcium concentration. The increase in burst duration is a consequence of the decrease in the calcium-activated potassium current (IK−Cin Eq. B.2a, also known as BK). Therefore, any decrease in this current by direct action of volatile anesthetics, 73–76should lead to additional burst duration. Because the increase in burst frequency is tied to reductions in the afterhyperpolarization current (IK-AHPin Eq. B.2a, also known as SK), it can be seen that if volatile anesthetics had the additional effect of reducing the afterhyperpolarization current, then burst frequency might be further increased.

Again, these results demonstrate that relatively modest volatile anesthetic effects on voltage-gated ion channels could have important consequences for the behavior of the neuron. The prolongation of burst duration in the presence of volatile anesthetics is known to occur in vitro . 93A switch from a bursting mode to a tonic-spiking mode for isolated hippocampal neurons in the presence of volatile anesthetics is not as well established experimentally, 94where network considerations may predominate. However, in simpler systems, a sufficiently high Ca2+conductance is necessary for bursting behavior to supplant spiking behavior. 95The ability of a volatile anesthetic to switch a neuron from a bursting pattern to a tonic spiking one whose rate is a function of the input current could have important consequences for information transfer in the brain. Switching of neural firing patterns is already used by the brain to regulate information flow through the thalamus where modulation of a T-type Ca2+channel determines the mode of behavior, 96and bursting is more likely to elicit cortical activity. 97 

As with the Morris–Lecar model, the relationship between specific effects in the model and a given anesthetic concentration is imprecise for many of the same reasons. An additional issue with the model of the CA3 hippocampal neuron used here is the presence of only a single type of Ca2+channel, when multiple HVA and LVA Ca2+channels contribute to the membrane potential of CA3 neurons, 58each with a unique sensitivity to the volatile anesthetics. 63,70–72Ultimately, the full impact of the effects of volatile anesthetics on the model neurons may not be realized until they are incorporated in a network which also incorporates the volatile anesthetic effects on inhibitory interactions 64as well as the contributions to network dynamics from other inhibitory and excitatory interactions which may not be affected by the volatile anesthetics.

Thalamic Reticular Nucleus.

An appreciation for the contribution of network interactions can be seen in the simulation of the Golomb–Rinzel model of the reticular nucleus of the thalamus. Here, a decrease in the rate at which the GABAAchannels close, which leads to prolongation of the inhibitory current, was used to incorporate the effects of intravenous or volatile anesthetic effects on inhibitory synaptic interaction. The anesthetic-induced reductions in the rate of GABAAchannel closure (βsynin Eq. C.3) that are introduced in the model correspond to those found experimentally for clinically relevant concentrations of anesthetics. 64,78As the open time of the GABAAchannels increased, progressive slowing of overall network activity, but not that of individual neurons, was observed, reflecting a growing synchrony in network behavior. Although there is a level of conductance for the GABAAchannel (ḡsynin Eq. C.1) below which synchrony becomes impossible, 53for the most part decreases in this conductance only altered the range of the rate parameter governing closure of the GABAAchannels (βsynin Eq. C.3) over which the transition to synchronous behavior is observed. As emphasized above for the model of the CA3 hippocampal neuron, volatile anesthetics may also affect the currents of the voltage-gated ion channels, particularly Ca2+channels. Although the model of the reticular nucleus used here incorporates a T-type Ca2+channel, many model features that might permit more elaborate interactions of volatile anesthetics with the individual neurons, particularly the Ca2+-dependent potassium currents were not present. Despite this, decreases in Ca2+current increased the ability to achieve network synchrony because this permitted a relatively greater influence of the inhibitory synaptic currents on membrane potential. Increasing the rate of GABAAchannel opening, as might occur with the benzodiazepines, did not affect model behavior since the time course of the opening was already quite fast relative to the time course of the membrane potential.

Additional aspects of anesthetic action on the behavior of the reticular nucleus of the thalamus could emerge with a more realistic model. Excitatory input from thalamocortical neurons, which are themselves inhibited by the reticular nucleus, is thought to contribute to regulation of the rhythm of the reticular nucleus. 48In addition, both  GABAAand GABABinhibitory interactions are present within the reticular nucleus and the inhibitory pathway from the reticular nucleus to the thalamocortical network. Moreover, the oscillations of both the thalamocortical neurons 98and the neurons of the reticular nucleus 99are driven in part by T-type Ca2+-currents which are known to be affected by volatile anesthetics, 70and whose modulation increased the synchrony of the model of the reticular nucleus, as shown above. Although GABABreceptors are not affected by volatile or intravenous anesthetics, 100the longer time constants associated with their activation and inactivation 101could impact on model behavior. When thalamocortical neurons and GABABinhibitory interactions are incorporated in models of the reticular nucleus, 59,60,102they appear to confer additional stability to the resulting oscillations. 59 

Fast-spiking Interneuron Network.

The Wang–Buzsáki model of a fast-spiking interneuron network, when stimulated with a level of input current sufficient to produce oscillations in the γ range (40–80 Hz), is notable for its paradoxical behavior when contrasted with the behavior of the reticular nucleus of the thalamus. Decreases in the rate constant governing closure of the GABAAchannels (βsynin Eq. D.4), as occurs with many intravenous and volatile anesthetics, lead to a loss  of synchrony (fig. 7), where the theoretical 103and computational 54basis for this behavior have been further elucidated elsewhere. Increases in the rate constant governing opening of GABAAchannels (αsynin Eq. D.4), which occurs with the benzodiazepines, may be more important to the behavior of this network (fig. 8) than in the relatively slower thalamic network. A much more complex model of γ oscillations in a hippocampal slice preparation, 8which incorporates the interaction of detailed multicompartment models of hippocampal neurons modulated by fast-spiking interneurons, requires more than a decrease in the rate of GABAAchannel closing to lose the synchrony of the γ frequency oscillations. An additional GABAAinhibitory leak current, as is documented to occur at higher concentrations of the intravenous anesthetics, 80,81,85,86must be incorporated to induce the loss of synchrony of the γ frequency oscillations. The addition of gap junctions to permit electrical coupling between the dendrites of interneurons is a model enhancement that can affect synchrony in certain types of networks. 104 

Implications for Anesthetic Action at the Ion Channel

The specific molecular site(s) responsible for general anesthesia remain(s) controversial. Since many of the intravenous and volatile anesthetics are known to exert their effect at the GABAAreceptor, it is tempting to implicate these actions at the GABAAreceptor as the unitary mechanism of general anesthesia. However, the volatile and intravenous anesthetics also affect voltage-gated ion channel activity, though it has been postulated that the anesthetic concentration at which these effects occur and the nature of these effects render them clinically irrelevant. 10The results presented above from the simulations with the Morris–Lecar model of the barnacle giant muscle fiber and the Pinsky–Rinzel model of the CA3 hippocampal neuron call this view into question. In each case, relatively modest reduction of Ca2+current corresponding to clinically relevant anesthetic concentrations brought about dramatic alterations in model behavior. Nevertheless, fundamental changes in network synchrony were observed when clinically relevant concentrations of intravenous or volatile anesthetics decreased the rate of GABAAchannel closure, and network synchrony could be further modulated by concurrent effects at the voltage gated channels.

The behavior of all four models is summarized in figure 9, where the concentration-effect curves for halothane on the voltage- and ligand-gated ion channels are shown along with the halothane concentrations at which model behavior is fundamentally altered. Although relatively high concentrations of halothane are necessary to achieve network synchrony in the thalamic model, this may be consistent with known differences of halothane on the electroencephalograph compared with other volatile anesthetics. For example, if it is assumed that the maximal reduction in the rate of GABAAchannel closing is twice as great for isoflurane as for halothane, 64,91then network synchrony is achieved for isoflurane concentrations of 0.40 mm (as opposed to 0.70 mm) in the thalamic model, and synchrony is lost for isoflurane concentrations of 0.16 mm (as opposed to 0.26 mm) in the fast spiking interneuron network.

Fig. 9. Concentration-effect curves of multiple ion channels in the presence of halothane together with anesthetic concentrations at which model behavior qualitatively changes for each of the four models described in the text. Arrows depict the halothane concentration at which quiescence occurs in the barnacle muscle fiber model (a), bursting is supplanted by spiking in the hippocampal model (b), synchrony is achieved in the network model of the reticular nucleus of the thalamus (c), and synchrony is lost in the fast-spiking interneuron model (d). The inhibition of a high voltage activated Ca2+channel, a Na+channel, and a K+channel in the presence of halothane are depicted using the Hill equation (Eq. 7) using values for the EC50of 0.85 mm, 632.0 mm, 66and 6.4 mm, 66respectively, and a Hill coefficient of 1.5. 10The prolongation of the time constant governing closure of GABAAchannels in the presence of halothane is depicted using the Hill equation assuming an EC50of 0.90 mm, a Hill coefficient of 1.5, and a maximal prolongation of 2.5 the baseline value. 64,91 

Fig. 9. Concentration-effect curves of multiple ion channels in the presence of halothane together with anesthetic concentrations at which model behavior qualitatively changes for each of the four models described in the text. Arrows depict the halothane concentration at which quiescence occurs in the barnacle muscle fiber model (a), bursting is supplanted by spiking in the hippocampal model (b), synchrony is achieved in the network model of the reticular nucleus of the thalamus (c), and synchrony is lost in the fast-spiking interneuron model (d). The inhibition of a high voltage activated Ca2+channel, a Na+channel, and a K+channel in the presence of halothane are depicted using the Hill equation (Eq. 7) using values for the EC50of 0.85 mm, 632.0 mm, 66and 6.4 mm, 66respectively, and a Hill coefficient of 1.5. 10The prolongation of the time constant governing closure of GABAAchannels in the presence of halothane is depicted using the Hill equation assuming an EC50of 0.90 mm, a Hill coefficient of 1.5, and a maximal prolongation of 2.5 the baseline value. 64,91 

Close modal

Implications as to the Mechanism of General Anesthesia

General anesthesia is not a single dichotomous state, and includes sedation, amnesia, loss of consciousness, immobility, and blunting of autonomic reflexes, with each of these features appearing once anesthetic concentrations reach a given threshold. 18–22Threshold behavior appears in the models presented here as well as in a more abstract model of general anesthetic action, 7and was the criterion used to generate figure 9. Clinically and encephalographically, 25general anesthesia can be biphasic, with an initial excitement phase followed by increasing depth of anesthesia as anesthetic concentration is increased. Several of the models presented here exhibit behavior consistent with inhibition and several models exhibit behavior consistent with excitement. In the CA3 hippocampal model, the increase in burst frequency, increase in burst duration and, perhaps, the shift to a tonically spiking pattern of behavior as anesthetic concentration is increased could be considered examples of increased excitability. The loss of synchrony in the fast-spiking interneuron model as GABAAchannels remain open longer could lead to less organized and, therefore, higher frequency electroencephalographic activity. Given the hypothesized role of this network in attention 41,42and in “binding” information from multiple cortical areas, 37it is conceivable that loss of synchrony in this network could be associated with sedation, amnesia, and decreased levels of consciousness. In the model of the reticular nucleus of the thalamus, which is known to be associated with sleep and barbiturate anesthesia, 43,47oscillations become more  organized as GABAAchannels remain open longer. Smaller decrements in the closing rate of the GABAAchannels are necessary to desynchronize the fast-spiking interneuronal network, whereas larger decrements are necessary to achieve complete synchrony in the model of the reticular nucleus of the thalamus. Thus, different components of the central nervous system may contribute to the overall state, and these contributions, even if involving the same receptor system, may appear at different anesthetic concentrations.

It may be important to appreciate that all aspects of anesthetic action on ion channel behavior and the subsequent impact on individual neurons and the networks they comprise may not be directly related to aspects of the anesthetic state that are clinically recognized and are of direct clinical utility. There is a growing appreciation for the dynamic qualities of the central nervous system, where neurons are continually adjusting their response in accordance with the prevailing pattern of pre- and postsynaptic activity. In various systems, this manifests itself in the form of synaptic plasticity, 105adjustments of dynamic range, 106and alterations in the conductances of populations of ion channels. 107It is conceivable that general anesthesia could affect these processes by altering the pattern of prevailing neural activity and through the dynamic adjustments in response to that activity. These changes could underlie some of the longer-term effects that have been reported following general anesthesia, 108and represent additional features of systems level behavior that may be amenable to computational modeling.

In summary, computational approaches offer additional tools for considering the integrated neural response to anesthetic action, and have demonstrated how anesthetic modulation of ion channel activity could lead to more complex systems level behavior. Collectively, the models described here have exhibited behavior consistent with both excitation and inhibition once anesthetic concentration reaches a clinically relevant threshold. Future progress in this area will require the development of more detailed biophysical models that more realistically incorporate the action of anesthetics and other drugs that modulate anesthetic action. Apart from elucidating the process by which one or more actions at the ion channel lead to general anesthesia, these approaches may eventually pave the way for more rational drug design and use.

The authors thank Marcel Poisit (Research Assistant, Krieger School of Arts and Sciences, Johns Hopkins University, Baltimore, MD) for his assistance with some of the preliminary simulations, and Drs. Roderick G. Eckenhoff (Associate Professor of Anesthesiology, University of Pennsylvania, Philadelphia, PA), Jonas Johansson (Assistant Professor of Anesthesiology, University of Pennsylvania, Philadelphia, PA), and Eric J. Moody (Associate Professor of Anesthesiology and Critical Care Medicine, Johns Hopkins Medical Institutions, Baltimore, MD) for their many suggestions.

Appendix A

Morris-Lecar model of the barnacle giant muscle fiber modified to permit volatile anesthetic modulation of Ca2+conductance. 49,56 

formula
formula
formula
formula
formula
formula
formula

TABLE 

Table. Model Parameters :

Table. Model Parameters :
Table. Model Parameters :

C = 20 μF/cm2, λ̄m= 1.0, λ̄m= 0.1

Initial Conditions :

V(0) =−20 mV, m(0) = 0.065, n(0) = 0.002

Definition of Model Parameters :

I = applied current (μA/cm2)

Ca, ḡK, ḡL= maximal conductance for Ca2+, K+, and leak currents (mS/cm2)

V = membrane potential (mV)

VCa, VK, VL= equilibrium potentials for Ca2+, K+, and leak currents (mV)

m, n = fraction of open Ca2+and K+channels

m(V), n(V) = fraction of open Ca2+and K+channels in steady state for given membrane potential

λm(V), λn(V) = rate constant for opening of Ca2+and K+channels (s−1)

λ̄m, λ̄n= maximal rate constant for opening of Ca2+and K+channels (s−1)

V1, V3= membrane potential at which m(V) and n(V) are equal to 0.5 (mV)

V2, V4= reciprocal slope of m(V) and n(V) (mV)

C = membrane capacitance (μF/cm2)

a = anesthetic concentration (mM)

a0.5−Ca= anesthetic concentration at which Ca2+-conductance is halved

nCa= Hill coefficient for anesthetic modulation of Ca2+-conductance

Appendix B

Pinsky-Rinzel Model of hippocampal CA-3 Neuron. 50,51 

formula
formula
formula
formula
formula
formula
formula
formula
formula
formula
formula
formula
formula
formula
formula
formula
formula

TABLE 

Table. No caption available.

Table. No caption available.
Table. No caption available.

Definition of Model Parameters :

VS, VD= membrane potentials of somatic and dendritic compartments (mV)

VNa, VK, VCa, VL= equilibrium potentials of sodium, potassium, calcium and leak current (mV)

IS, ID= applied current to somatic and dendritic compartments (μA/cm2)

INa, IK−DR, IK−C, IK−AHP, ICa, IL= sodium, potassium (delayed rectifier), calcium dependent potassium, potassium afterhyperpolarization, calcium, and leak currents (μA/cm2)

m, h = gating variables for activation (m) and inactivation (h) of INa

n, s, c, q = gating variables for activation of IK−DR, ICa, IK−C, IK−AHP

αm, βm, αh, βh, αn, βn, αs, βs, αc, βc, αq, βq= kinetic parameters governing activity of the gating variables

m= equilibrium value for gating variable m obtained per Eq. B.9c

χ= saturating function of calcium concentration

Na, ḡCa, ḡK−DR, ḡK−C, ḡK−AHP, ḡL= maximal conductance of sodium, calcium, delayed rectifier, calcium activated potassium, potassium afterhyperpolarization, and leak currents (mS/cm2)

C = membrane capacitance (μF/cm2)

gC= coupling strength between somatic and dendritic compartment (mS/cm2)

ρ= fraction of total neuron area in somatic compartment

Appendix C

Golomb-Rinzel model of reticular nucleus of thalamus. 53 

formula
formula
formula
formula

Model Parameters :

VCa= 120 mV, VL=−60 mV, Vsyn=−80 mV, ḡCa= 0.5 mS/cm2, ḡL= 0.05 mS/cm2, ḡsyn= 0.38 mS/cm2, αsyn= 1 ms−1, βsyn= .08 ms−1, C = 1 μF/cm2, φ= 2 ms−1, N = 100, y = .8

Initial Conditions :

Vi(0) randomly and uniformly distributed between −90 mV and −50 mV (hyperpolarized initial conditions) or −50 mV and −10 mV (depolarized initial conditions). si(0) and hi(0) obtained from steady state solutions of Eqs. C.2 and C.3for the corresponding value of Vi(0).

Definition of Model Parameters :

Vi= membrane potential of i-th neuron (mV)

N = number of neurons in network

γ= fraction of total number of neurons connected to an individual neuron

Sij= connection matrix of either zeros or ones, with zeros on the diagonal, where connections were generated randomly so that, on average, γN connections were made to each neuron

VCa, VL, Vsyn= equilibrium potential of calcium, leak, and synaptic currents (mV)

Ca, ḡL, ḡsyn= maximal conductance of calcium, leak, and synaptic currents (mS/cm2)

hi, si= gating variable for calcium and synaptic currents for the i-th neuron, representing the fraction of channels that are open

αsyn, βsyn= forward and backward rate constants controlling dynamics of gating variable siwhich represents the fraction of open GABAAchannels (ms−1)

φ= parameter governing time constant of gating variable for calcium channel

C = membrane capacitance (μF/cm2)

Appendix D

Wang-Buzsáki model of hippocampal interneuronal network. 54 

formula
formula
formula
formula
formula
formula
formula
formula

Model Parameters :

VNa= 55 mV, VK=−90 mV, VL=−65 mV, Vsyn=−75 mV, ḡNa= 35 mS/cm2, ḡK= 9 mS/cm2, ḡL= 0.1 mS/cm2, ḡsyn= 0.1 mS/cm2, φ= 5, Θsyn= 0 mV, C = 1 μF/cm2, αsyn= 12 ms−1, βsyn= 0.08 ms−1, γ= 0.6, N = 100, Ii= injection current to i-th neuron (chosen from a Gaussian distribution with mean 1.0 μA/cm2and standard deviation .03 μA/cm2)

Initial Conditions :

Vi(0) randomly and uniformly distributed between −70 mV and −50 mV. hi(0), ni(0), and si(0) obtained from steady state solutions of Eqs. D.2–4for the corresponding value of Vi(0).

Definition of Model Parameters :

Vi= membrane potential of i-th neuron (mV)

N = number of neurons in network

γ= fraction of total number of neurons connected to an individual neuron

Sij= connection matrix of either zeros or ones, with zeros on the diagonal, where connections were generated randomly so that, on average, γN connections were made to each neuron

VNa, VK, VL, Vsyn= equilibrium potential of sodium, potassium, leak, and synaptic currents (mV)

Na, ḡK, ḡL, ḡsyn= maximal conductance of sodium, potassium, leak, and synaptic currents (mS/cm2)

mi, hi, ni, si= gating variables governing activation and inactivation of sodium current, activation of potassium current, and activation of inhibitory GABAAcurrent

m= steady state value of gating variable governing activation of sodium current

φ= dimensionless parameter governing time constant of gating variable governing sodium inactivation and potassium activation

αh, βh, αn, βn= kinetic parameters governing activity of the gating variables hi, and ni

αsyn, βsyn= forward and backward rate constants controlling dynamics of gating variable siwhich represents the fraction of open GABAAchannels (ms−1)

Ii= injection current to i-th neuron (μA/cm2). Injection currents, although constant throughout the simulation for each neuron, were not uniform and were chosen from a Gaussian distribution.

F(Vj) = modulates synaptic opening as a function of presynaptic membrane potential

Θ= presynaptic membrane potential at which synaptic opening rate reaches its half-maximal value

C = membrane capacitance (μF/cm2)

1.
Yaksh TL, Lynch C, III, Zapol WM, Biebuyck JF, Saidman LJ (eds.): Anesthesia Biologic Foundations. Philadelphia, Lippincott-Raven, 1998, pp 5–286
2.
Eckenhoff RG, Johansson JS: On the relevance of “clinically relevant concentrations” of inhaled anesthetics in in vitro experiments. A nesthesiology 1999; 91: 856–60
3.
Eger EI, Fisher DM, Dilger JP, Sonner JM, Evers A, Franks NP, Harris RA, Kendig JJ, Lieb WR, Yamakura T: Relevant concentrations of inhaled anesthetics for in vitro studies of anesthetic mechanisms. A nesthesiology 2001; 94: 915–21
4.
Gottschalk A: Seeking an integrated model of anesthetic action. A nesthesiology 2001; 95: 1536–7
5.
Koch C, Segev I, (eds.): Methods in Neuronal Modeling From Ions to Networks, 2nd edition. Cambridge, MA, MIT Press, 1998
6.
Guckenheimer J, Holmes P: Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vector Fields, 2nd edition. New York, NY, Springer-Verlag, 1983
7.
Steyn-Ross ML, Steyn-Ross DA, Sleigh JW, Liley DTJ: Theoretical electroencephalogram stationary spectrum for a white-noise-driven cortex: Evidence for a general anesthetic-induced phase transition. Physical Review E 1999; 60: 7299–311
8.
Faulkner HJ, Traub RD, Whittington MA: Disruption of synchronous gamma oscillations in the rat hippocampal slice: a common mechanism of anaesthetic drug action. Br J Pharmacol 1998; 125: 483–92
9.
Faulkner HJ, Traub RD, Whittington MA: Anaesthetic/amnesic agents disrupt beta frequency oscillations associated with potentiation of excitatory synaptic potentials in the rat hippocampal slice. Br J Pharmacol 1999; 128: 1813–25
10.
Franks NP, Lieb WR: Molecular and cellular mechanisms of general anaesthesia. Nature 1994; 367: 607–14
11.
Carlson B, Hales T, Olsen R.: GABAAReceptors and Anesthesia, Anesthesia: Biologic Foundations. Edited by Yaksh TL, Lynch C, III, Zapol WM, Biebuyck JF, Saidman LJ. Philadelphia, Lippincott-Raven Publishers, 1997, pp 259–71
12.
Quinlan JJ, Homanics GE, Firestone LL: Anesthesia sensitivity in mice that lack the beta3 subunit of the gamma-aminobutyric acid type A receptor. A nesthesiology 1998; 88: 775–80
13.
Homanics GE, Ferguson C, Quinlan JJ, Daggett J, Snyder K, Lagenaur C, Mi ZP, Wang XH, Grayson DR, Firestone LL: Gene knockout of the alpha6 subunit of the gamma-aminobutyric acid type A receptor: lack of effect on responses to ethanol, pentobarbital, and general anesthetics. Mol Pharmacol 1997; 51: 588–96
14.
Raines DE, Claycomb RJ, Scheller M, Forman SA: Nonhalogenated alkane anesthetics fail to potentiate agonist actions on two ligand-gated ion channels. A nesthesiology 2001; 95: 470–7
15.
de Sousa SL, Dickinson R, Lieb WR, Franks NP: Contrasting synaptic actions of the inhalational general anesthetics isoflurane and xenon. A nesthesiology 2000; 92: 1055–66
16.
Yamakura T, Harris RA: Effects of gaseous anesthetics nitrous oxide and xenon on ligand-gated ion channels. Comparison with isoflurane and ethanol. A nesthesiology 2000; 93: 1095–101
17.
Eger EI, Saidman LJ, Brandstater B: Minimum alveolar anesthetic concentration: a standard of anesthetic potency. A nesthesiology 1965; 26: 756–63
18.
Stoelting RK, Longnecker DE, Eger EI: Minimum alveolar concentrations in man on awakening from methoxyflurane, halothane, ether and fluroxene anesthesia: MAC awake. A nesthesiology 1970; 33: 5–9
19.
Yakaitis RW, Blitt CD, Angiulo JP: End-tidal halothane concentration for endotracheal intubation. A nesthesiology 1977; 47: 386–8
20.
Roizen MF, Horrigan RW, Frazer BM: Anesthetic doses blocking adrenergic (stress) and cardiovascular responses to incision–MAC BAR. A nesthesiology 1981; 54: 390–8
21.
de Jong RH, Eger EI: MAC expanded: AD50 and AD95 values of common inhalation anesthetics in man. A nesthesiology 1975; 42: 384–9
22.
Veselis RA, Reinsel RA, Feshchenko VA: Drug-induced amnesia is a separate phenomenon from sedation: electrophysiologic evidence. A nesthesiology 2001; 95: 896–907
23.
Antognini JF, Schwartz K: Exaggerated anesthetic requirements in the preferentially anesthetized brain. A nesthesiology 1993; 79: 1244–9
24.
Borges M, Antognini JF: Does the brain influence somatic responses to noxious stimuli during isoflurane anesthesia? A nesthesiology 1994; 81: 1511–5
25.
Levy WJ: Neurophysiologic brain monitoring: electroencephalography, Anesthesia and Neurosurgery, 4th edition. Edited by Cottrell JE, Smith DS. Philadelphia, PA, Mosby, 2001, pp 201–17
26.
Rampil IJ, Laster MJ: No correlation between quantitative electroencephalographic measurements and movement response to noxious stimuli during isoflurane anesthesia in rats. A nesthesiology 1992; 77: 920–5
27.
Johansen JW, Sebel PS: Development and clinical application of electroencephalographic bispectrum monitoring. A nesthesiology 2000; 93: 1336–44
28.
Pichlmayr I, Lips U:[EEG effects of diazepam. Studies on 16 patients (author's transl)]. Anaesthesist 1980; 29: 317–27
29.
Rampil IJ, Lockhart SH, Eger EI, Yasuda N, Weiskopf RB, Cahalan MK: The electroencephalographic effects of desflurane in humans. A nesthesiology 1991; 74: 434–9
30.
Clark DL, Hosick EC, Rosner BS: Neurophysiological effects of different anesthetics in unconscious man. J Appl Physiol 1971; 31: 884–91
31.
Brackman LE, Lümofstrümom B, Widen L: Electroencephalography in halothane anaesthesia. Acta Anaesthesiol Scand 1964; 8: 115–30
32.
Ehlers CL, Havstad J, Prichard D, Theiler J: Low doses of ethanol reduce evidence for nonlinear structure in brain activity. J Neurosci 1998; 18: 7474–86
33.
Heinke W, Schwarzbauer C: Subanesthetic isoflurane affects task-induced brain activation in a highly specific manner: a functional magnetic resonance imaging study. A nesthesiology 2001; 94: 973–81
34.
Alkire MT, Pomfrett CJ, Haier RJ, Gianzero MV, Chan CM, Jacobsen BP, Fallon JH: Functional brain imaging during anesthesia in humans: effects of halothane on global and regional cerebral glucose metabolism. A nesthesiology 1999; 90: 701–9
35.
Alkire MT, Haier RJ, Fallon JH: Toward a unified theory of narcosis: brain imaging evidence for a thalamocortical switch as the neurophysiologic basis of anesthetic-induced unconsciousness. Conscious Cogn 2000; 9: 370–86
36.
Alkire MT: Quantitative EEG correlations with brain glucose metabolic rate during anesthesia in volunteers. A nesthesiology 1998; 89: 323–33
37.
Singer W: Neuronal synchrony: a versatile code for the definition of relations? Neuron 1999; 24: 49–25
38.
Jefferys JG, Traub RD, Whittington MA: Neuronal networks for induced ’40 Hz’ rhythms. Trends Neurosci 1996; 19: 202–8
39.
Galarreta M, Hestrin S: A network of fast-spiking cells in the neocortex connected by electrical synapses. Nature 1999; 402: 72–5
40.
Galarreta M, Hestrin S: Spike transmission and synchrony detection in networks of GABAergic interneurons. Science 2001; 292: 2295–9
41.
Vaadia E, Haalman I, Abeles M, Bergman H, Prut Y, Slovin H, Aertsen A: Dynamics of neuronal interactions in monkey cortex in relation to behavioural events. Nature 1995; 373: 515–8
42.
Steinmetz PN, Roy A, Fitzgerald PJ, Hsiao SS, Johnson KO, Niebur E: Attention modulates synchronized neuronal firing in primate somatosensory cortex. Nature 2000; 404: 187–90
43.
Steriade M, Deschenes M: The thalamus as a neuronal oscillator. Brain Res 1984; 320: 1–63
44.
Newman J: Putting the puzzle together. Part I: Towards a general theory of the neural correlates of consciousness. J Consciousness Stud 1997; 4: 46–66
45.
Lumer ED, Edelman GM, Tononi G: Neural dynamics in a model of the thalamocortical system. I. Layers, loops and the emergence of fast synchronous rhythms. Cereb Cortex 1997; 7: 207–27
46.
Llinas R, Ribary U, Contreras D, Pedroarena C: The neuronal basis for consciousness. Philos Trans R Soc Lond B Biol Sci 1998; 353: 1841–9
47.
Deschenes M, Paradis M, Roy JP, Steriade M: Electrophysiology of neurons of lateral thalamic nuclei in cat: resting properties and burst discharges. J Neurophysiol 1984; 51: 1196–219
48.
Steriade M: Cellular substrates of brain rhythms, Electroencephalography: Basic Principles, Clinical Applications, and Related Fields, 3rd Edition. Edited by Niedermeyer E, da Silva FL. Baltimore, Williams & Wilkins, 1993, pp 27–62
49.
Morris C, Lecar H: Voltage oscillations in the barnacle giant muscle fiber. Biophys J 1981; 35: 193–213
50.
Pinsky PF, Rinzel J: Intrinsic and network rhythmogenesis in a reduced Traub model for CA3 neurons. J Comput Neurosci 1994; 1: 39–60
51.
Pinsky PF, Rinzel J: Erratum. J Comput Neurosci 1995; 2: 275
52.
Traub RD, Wong RK, Miles R, Michelson H: A model of a CA3 hippocampal pyramidal neuron incorporating voltage-clamp data on intrinsic conductances. J Neurophysiol 1991; 66: 635–50
53.
Golomb D, Rinzel J: Clustering in globally coupled inhibitory neurons. Physica D 1994; 72: 259–82
54.
Wang XJ, Buzsaki G: Gamma oscillation by synaptic inhibition in a hippocampal interneuronal network model. J Neurosci 1996; 16: 6402–13
55.
Hodgkin AL, Huxley AF: A quantitative description of membrane current and its application to conduction and excitation in nerve. J Physiol 1952; 117: 500–44
56.
Rinzel J, Ermentrout B: Analysis of neural excitability and oscillations, Methods in Neuronal Modeling, 2nd Edition. Edited by Koch C, Segev I. Cambridge, MA, The MIT Press, 1998, pp 251–91
57.
Traub RD, Jefferys JG, Miles R, Whittington MA, Toth K: A branching dendritic model of a rodent CA3 pyramidal neurone. J Physiol 1994; 481 ( Pt 1): 79–95
58.
Migliore M, Cook EP, Jaffe DB, Turner DA, Johnston D: Computer simulations of morphologically reconstructed CA3 hippocampal neurons. J Neurophysiol 1995; 73: 1157–68
59.
Golomb D, Wang XJ, Rinzel J: Synchronization properties of spindle oscillations in a thalamic reticular nucleus model. J Neurophysiol 1994; 72: 1109–26
60.
Destexhe A, Bal T, McCormick DA, Sejnowski TJ: Ionic mechanisms underlying synchronized oscillations and propagating waves in a model of ferret thalamic slices. J Neurophysiol 1996; 76: 2049–70
61.
Buzsaki G, Leung LW, Vanderwolf CH: Cellular bases of hippocampal EEG in the behaving rat. Brain Res 1983; 287: 139–71
62.
Pike FG, Goddard RS, Suckling JM, Ganter P, Kasthuri N, Paulsen O: Distinct frequency preferences of different types of rat hippocampal neurones in response to oscillatory input currents. J Physiol 2000; 529 Pt 1: 205–13
63.
Herrington J, Stern RC, Evers AS, Lingle CJ: Halothane inhibits two components of calcium current in clonal (GH3) pituitary cells. J Neurosci 1991; 11: 2226–40
64.
Banks MI, Pearce RA: Dual actions of volatile anesthetics on GABA(A) IPSCs: dissociation of blocking and prolonging effects. A nesthesiology 1999; 90: 120–34
65.
Haydon DA, Urban BW: The effects of some inhalation anaesthetics on the sodium current of the squid giant axon. J Physiol 1983; 341: 429–39
66.
Haydon DA, Urban BW: The actions of some general anaesthetics on the potassium current of the squid giant axon. J Physiol 1986; 373: 311–27
67.
Patel AJ, Honore E, Lesage F, Fink M, Romey G, Lazdunski M: Inhalational anesthetics activate two-pore-domain background K+ channels. Nat Neurosci 1999; 2: 422–6
68.
Sirois JE, Lei Q, Talley EM, Lynch C, III, Bayliss DA: The TASK-1 two-pore domain K+ channel is a molecular substrate for neuronal effects of inhalation anesthetics. J Neurosci 2000; 20: 6347–54
69.
Patel AJ, Honore E: Anesthetic-sensitive 2P domain K+ channels. A nesthesiology 2001; 95: 1013–21
70.
McDowell TS, Pancrazio JJ, Barrett PQ, Lynch C, III: Volatile anesthetic sensitivity of T-type calcium currents in various cell types. Anesth Analg 1999; 88: 168–73
71.
Study RE: Isoflurane inhibits multiple voltage-gated calcium currents in hippocampal pyramidal neurons. A nesthesiology 1994; 81: 104–16
72.
Kamatchi GL, Chan CK, Snutch T, Durieux ME, Lynch C, III: Volatile anesthetic inhibition of neuronal Ca channel currents expressed in Xenopus oocytes. Brain Res 1999; 831: 85–96
73.
Tas PW, Kress HG, Koschel K: Volatile anesthetics inhibit the ion flux through Ca2+-activated K+ channels of rat glioma C6 cells. Biochim Biophys Acta 1989; 983: 264–8
74.
Pancrazio JJ, Park WK, Lynch C, III: Effects of enflurane on the voltage-gated membrane currents of bovine adrenal chromaffin cells. Neurosci Lett 1992; 146: 147–51
75.
Pancrazio JJ, Park WK, Lynch C, III: Inhalational anesthetic actions on voltage-gated ion currents of bovine adrenal chromaffin cells. Mol Pharmacol 1993; 43: 783–94
76.
Namba T, Ishii TM, Ikeda M, Hisano T, Itoh T, Hirota K, Adelman JP, Fukuda K: Inhibition of the human intermediate conductance Ca(2+)-activated K(+) channel, hIK1, by volatile anesthetics. Eur J Pharmacol 2000; 395: 95–101
77.
Macdonald RL, Twyman RE, Rogers CJ, Weddle MG: Pentobarbital regulation of the kinetic properties of GABA receptor chloride channels. Adv Biochem Psychopharmacol 1988; 45: 61–71
78.
Whittington MA, Jefferys JG, Traub RD: Effects of intravenous anaesthetic agents on fast inhibitory oscillations in the rat hippocampus in vitro. Br J Pharmacol 1996; 118: 1977–86
79.
Proctor WR, Mynlieff M, Dunwiddie TV: Facilitatory action of etomidate and pentobarbital on recurrent inhibition in rat hippocampal pyramidal neurons. J Neurosci 1986; 6: 3161–8
80.
Hales TG, Lambert JJ: The actions of propofol on inhibitory amino acid receptors of bovine adrenomedullary chromaffin cells and rodent central neurones. Br J Pharmacol 1991; 104: 619–28
81.
Barker JL, Harrison NL, Lange GD, Owen DG: Potentiation of gamma-aminobutyric-acid-activated chloride conductance by a steroid anaesthetic in cultured rat spinal neurones. J Physiol 1987; 386: 485–501
82.
Twyman RE, Macdonald RL: Neurosteroid regulation of GABAA receptor single-channel kinetic properties of mouse spinal cord neurons in culture. J Physiol 1992; 456: 215–45
83.
Birnir B, Tierney ML, Dalziel JE, Cox GB, Gage PW: A structural determinant of desensitization and allosteric regulation by pentobarbitone of the GABAA receptor. J Membr Biol 1997; 155: 157–66
84.
Zhu WL, Vicini S: Neurosteroid prolongs GABAA channel deactivation by altering kinetics of desensitized states. J Neurosci 1997; 17: 4022–31
85.
Evans RH, Hill RG: GABA-mimetic action of etomidate. Experientia 1978; 34: 1325–7
86.
Schulz DW, Macdonald RL: Barbiturate enhancement of GABA-mediated inhibition and activation of chloride ion conductance: correlation with anticonvulsant and anesthetic actions. Brain Res 1981; 209: 177–88
87.
Study RE, Barker JL: Diazepam and (–)-pentobarbital: fluctuation analysis reveals different mechanisms for potentiation of gamma-aminobutyric acid responses in cultured central neurons. Proc Natl Acad Sci USA 1981; 78: 7180–4
88.
Wolfram S: The Mathematica Book, 4th edition. New York, NY, Wolfram Media/Cambridge University Press, 1999
89.
Carnahan B, Luther HA, Wilkes JO: Applied Numerical Methods. New York, NY, John Wiley and Sons, 1969; pp 361–80
90.
Bower JM, Beeman D: The Book of GENESIS: Exploring Realistic Neural Models with the General Neural Simulation System, 2nd edition. New York, NY, Springer-Verlag, 1998
91.
Jones MV, Harrison NL: Effects of volatile anesthetics on the kinetics of inhibitory postsynaptic currents in cultured rat hippocampal neurons. J Neurophysiol 1993; 70: 1339–49
92.
Miao N, Nagao K, Lynch C, III: Thiopental and methohexital depress Ca2+ entry into and glutamate release from cultured neurons. A nesthesiology 1998; 88: 1643–53
93.
Pearce RA: Volatile anaesthetic enhancement of paired-pulse depression investigated in the rat hippocampus in vitro. J Physiol 1996; 492 ( Pt 3): 823–40
94.
Fujiwara N, Higashi H, Nishi S, Shimoji K, Sugita S, Yoshimura M: Changes in spontaneous firing patterns of rat hippocampal neurones induced by volatile anaesthetics. J Physiol 1988; 402: 155–75
95.
Turrigiano G, LeMasson G, Marder E: Selective regulation of current densities underlies spontaneous changes in the activity of cultured neurons. J Neurosci 1995; 15: 3640–52
96.
Sherman SM: Tonic and burst firing: dual modes of thalamocortical relay. Trends Neurosci 2001; 24: 122–6
97.
Swadlow HA, Gusev AG: The impact of ’bursting’ thalamic impulses at a neocortical synapse. Nat Neurosci 2001; 4: 402–8
98.
McCormick DA, Huguenard JR: A model of the electrophysiological properties of thalamocortical relay neurons. J Neurophysiol 1992; 68: 1384–400
99.
Destexhe A, Contreras D, Steriade M, Sejnowski TJ, Huguenard JR: In vivo, in vitro, and computational analysis of dendritic calcium currents in thalamic reticular neurons. J Neurosci 1996; 16: 169–85
100.
Bowery NG, Price GW, Hudson AL, Hill DR, Wilkin GP, Turnbull MJ: GABA receptor multiplicity. Visualization of different receptor types in the mammalian CNS. Neuropharmacology 1984; 23: 219–31
101.
Destexhe A, Mainen ZF, Sejnowski TJ: Synthesis of models for excitable membranes, synaptic transmission and neuromodulation using a common kinetic formalism. J Comput Neurosci 1994; 1: 195–230
102.
Wang XJ, Golomb D, Rinzel J: Emergent spindle oscillations and intermittent burst firing in a thalamic model: specific neuronal mechanisms. Proc Natl Acad Sci USA 1995; 92: 5577–81
103.
White JA, Chow CC, Ritt J, Soto-Trevino C, Kopell N: Synchronization and oscillatory dynamics in heterogeneous, mutually inhibited neurons. J Comput Neurosci 1998; 5: 5–16
104.
Traub RD, Kopell N, Bibbig A, Buhl EH, LeBeau FE, Whittington MA: Gap junctions between interneuron dendrites can enhance synchrony of gamma oscillations in distributed networks. J Neurosci 2001; 21: 9478–86
105.
Maren S, Baudry M: Properties and mechanisms of long-term synaptic plasticity in the mammalian brain: relationships to learning and memory. Neurobiol Learn Mem 1995; 63: 1–18
106.
Turrigiano GG, Leslie KR, Desai NS, Rutherford LC, Nelson SB: Activity-dependent scaling of quantal amplitude in neocortical neurons. Nature 1998; 391: 892–6
107.
Marder E, Abbott LF, Turrigiano GG, Liu Z, Golowasch J: Memory from the dynamics of intrinsic membrane currents. Proc Natl Acad Sci U.S.A 1996; 93: 13481–6
108.
Moller JT, Cluitmans P, Rasmussen LS, Houx P, Rasmussen H, Canet J, Rabbitt P, Jolles J, Larsen K, Hanning CD, Langeron O, Johnson T, Lauven PM, Kristensen PA, Biedler A, van Beem H, Fraidakis O, Silverstein JH, Beneken JE, Gravenstein JS: Long-term postoperative cognitive dysfunction in the elderly ISPOCD1 study. ISPOCD investigators. International Study of Post-Operative Cognitive Dysfunction. Lancet 1998; 351: 857–61